Вы находитесь на странице: 1из 12

Green Chemistry Highlights

pubs.acs.org/OPRD

The Green Chemistry Articles of Interest


1. INTRODUCTION An application of this new solvent in hydride reduction reac-
The American Chemical Society’s (ACS) Green Chemistry tions is reported, and a specific new process for the reduction
Institute (GCI) Pharmaceutical Roundtable (PR) was developed of nitriles to amines in 2-MeTHF and in 1,2,3-TMP in using
in 2005 to encourage the integration of green chemistry and 1,1,3,3-tetramethyldisiloxane (TMDS) in combination with
green engineering into the pharmaceutical industry. copper triflate Cu(OTf)2 is developed and described (Green
The Roundtable currently has 16 member companies as Chem. 2013, 15, 3020−3026).
compared to three in 2005. The membership scope has also McGonagle et al. have published a solvent selection guide
broadened to include contract research/manufacturing organ- for aldehyde-based direct reductive amination processes. A
izations, generic pharmaceuticals, and related companies. SciFinder survey indicated that 25% of published reductive
Members currently include ACS GCI, Amgen, AstraZeneca, aminations use either dichloromethane (DCM) or dichloroethane
Boehringer-Ingelheim, Bristol-Myers Squibb, Codexis, Dr. (DCE) as solvent. In the study three commonly used reagents are
Reddy’s, DSM Pharmaceutical Products, Eli Lilly and Company, examined, sodium cyanoborohydride, sodium triacetoxyborohy-
GlaxoSmithKline, Johnson & Johnson, Merck & Co., Inc., dride (STAB), and picoline−borane complexes. The results for
Novartis, Pfizer, Inc., Roche, and Sanofi. reactions with DCM and DCE are compared with greener
One of the strategic priorities of the Roundtable is to inform solvents such as ethyl acetate, IPA, and certain ethers. The results
and influence the research agenda. Two of the first steps to of the study show that chlorinated solvents are not required,
achieve this objective were to publish a paper outlining key green particularly when using STAB as the reducing agent (Green
chemistry research areas from a pharmaceutical perspective Chem. 2013, 15, 1159−1165).
(Green. Chem. 2007, 9, 411−420) and to establish annual ACS
GCIPR research grants. This document follows on from the 3. AMIDE FORMATION
Green Chemistry paper and is largely based on the key research Kang et al. have reported on the reaction of nitriles and alcohols
areas, though new sections have been added. The review period under ruthenium-catalyzed conditions to generate amides in a
covers April 2013−September 2013. redox-neutral single step with complete atom economy. The
These articles of interest represent the opinions of the authors reaction differentiates itself from previous amidation reactions
and do not necessarily represent the views of the member utilizing these components in that the carbon of the nitrile is not
companies. Some articles are included because, whilst not the carbonyl source in the reaction, and instead the reaction
currently being regarded as green, the chemistry has the proceeds through hydrogen transfer from the alcohol to nitrile
potential to improve the current state of the art if developed with the subsequent C−N bond formation between the nitrogen
further. The inclusion of an article in this document does not of the nitrile and the α-carbon of the primary alcohol. Screening
give any indication of safety or operability. Anyone wishing to studies on a model system indicated that RuH2(CO)(PPh3)3
use any reaction or reagent must consult and follow their with an NHC precursor (1,3-diisopropylimidazolium bromide)
internal chemical safety and hazard procedures. as ligand were the optimal catalyst system with toluene at reflux
as the solvent and sodium hydride added as base. A range of
2. SOLVENTS both aliphatic and aromatic nitriles were successful substrates
with the main limitation being functional groups that were
A special issue of Current Organic Chemistry was dedicated to unable to tolerate the basic reaction conditions. Similarly, good
green solvents in the synthesis of pharmaceuticals (Curr. Org. substrate scope was observed with respect to the primary
Chem. 2013, 17, 1131). Next to articles on biocatalysis and alcohol employed. Given the relative ease of accessing isotope-
dynamic kinetic resolution, a review paper on green solvents in labelled nitriles, this methodology was demonstrated to be
organic synthesis is included, covering several main classes of effective for the synthesis of labelled-amides with complete
environmentally benign solvents, namely water, fluorous solvents, isotope incorporation. Deuterium labelling at the α-CH2 of the
ionic liquids, organic carbonates, scCO2, and biosolvents. The alcohol demonstrated unequivocally that hydrogen generated
focus of the article is on new evolutions in the last 15 years and from alcohol oxidation is responsible for the reduction of the
on the complementarity of these alternatives (Curr. Org. Chem. nitrile, and NMR studies confirmed the involvement of imine
2013, 17, 1179−1187). intermediates (J. Am. Chem. Soc. 2013, 135, 11704−11707).
Liu et al. published a communication on catalysts and
additive-free synthesis of disulfides in the biobased green
solvent ethyl lactate. This solvent was compared to traditional
solvents like DMF and toluene and showed an interesting
promotion effect of ethyl lactate on this thiol-coupling reaction
(RSC Adv. 2013, 3, 21369−21372).
Sutter et al. evaluated the toxicological profile of the glycerol-
based 1,2,3-trimethoxypropane. It is proposed by the authors
to be an alternative green solvent as it gives a negative result in Received: May 27, 2014
the Ames test, has low acute toxicity, relatively high flash- Accepted: June 5, 2014
point (45 °C), and low ecotoxicity in an aquatic environment.

© XXXX American Chemical Society A dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

Sun et al. have reported on the formation of amides from long-chain monocarboxylic acids are not good substrates. A
aromatic amines utilizing 1,3-diketones as the acylating agent. plausible mechanism is provided in which the nano-MgO activates
The reaction proceeds via a novel C−C bond cleavage under the carboxylic acid on its surface throughout the amide bond
oxidative metal-free conditions with water as the solvent. Initial formation (Appl. Catal., A 2013, 456, 118−125).
screening indicated 30% hydrogen peroxide to be the optimal
oxidant. Various anilines were successful substrates for the
reaction though those bearing electron-withdrawing substitu-
ents required elevated reaction temperatures. Aliphatic amines
failed to react under the reaction conditions. Various diketones
were also evaluated as acylating agents, and steric hindrance
was demonstrated to be a critical factor in the efficiency of the Two groups have published on the oxidative amidation of
acyl transfer. Control experiments to determine the mechanism benzylic alcohols catalyzed by iron salts in the presence of an
suggest that the reaction proceeds through an enamine oxidant. Gaspa et al. have utilized FeCl3·6H2O with 70%
intermediate via a radical mechanism (Green Chem. 2013, 15, aqueous TBHP as the oxidant. Initially the amine is reacted
3289−3294). with NCS to generate the N-chloroamine in situ, which can
then be consecutively treated with the alcohol, catalyst, and
oxidant. Electronic or steric effects on the aromatic ring of the
benzylic alcohol were shown to have little influence on the
reaction, and both primary and secondary amines were well
tolerated. A radical-type mechanism is proposed (Org. Biomol.
Chem. 2013, 11, 3803−3807). The system of Ghosh et al.
utilizes Fe(NO3)3·9H2O as the catalyst, and the reaction
proceeds in a tandem fashion. Initially, the alcohol is oxidized in
Lanigan et al. have published on the use of B(OCH2CF3)3 to the presence of catalytic amounts of the iron salt and TEMPO
promote the reaction between carboxylic acids and amines with air as the stoichiometric oxidant. After 2 h, the amine
to generate the corresponding amides. An alternative scalable hydrochloride salt, calcium carbonate, and 70% aqueous TBHP
synthesis of the reagent from B2O3 is described, and the are added, and the reaction is heated to 60 °C for 16 h to form
compound can be stored at room temperature under nitrogen the amide. Again, a range of primary and secondary amines
for ∼4 months without decomposition. In addition, a novel were well tolerated, and little effect was noted based on the
work-up has been developed using a series of commercially electronic properties of the benzylic alcohol. In the case of
available scavenger resins to enable isolation of the pure amide chiral amine salts, no racemization was detected during the
directly from the reaction mixture. The scope of the amida- amidation, and again a radical-type mechanism is proposed
tion process is also evaluated. In examples with acids with (Tetrahedron Lett. 2013, 54, 4922−4925).
adjacent chiral centres, no significant racemization was involved,
and the protocol was extended to the formation of dipeptides.
B(OCH2CF3)3 was also demonstrated to be an effective reagent
for the formylation of both primary and secondary amines with
DMF (J. Org. Chem. 2013, 78, 4512−4523).

4. OXIDATIONS
C−H functionalization is an important area for green chemistry
Das et al. have published on the use of nano magnesium since it offers opportunities for shorter synthetic routes with
oxide (MgO) to promote the formation of amides under reduced waste. Roduner et al. reviewed recent advances in
solvent-free reaction conditions. The reactions took place at C−H oxidations using molecular oxygen, covering both organo-
70 °C with the catalyst being recovered by centrifugation, and metallic and biocatalytic approaches. The authors address a critical
able to be recycled through 5 reactions without loss of activity. aspect of reactions using oxygen that is not often addressed in
The optimum catalyst load was 5 mol %, with a loss of activity publications: safety. Two ways of safely using oxygen are
noted at higher loadings, which is believed to be due to a reviewed, including supercritical CO2 and water as solvents, and
decrease in surface area due to particle aggregation. The catalyst flow chemistry in microreactors (ChemCatChem 2013, 5,
was characterized by numerous techniques such as EDX, FTIR 82−112).
and SEM micrographs, and changes observed as the catalyst Use of biocatalysis for oxidations is an area of growing
lost activity. The activity of the catalyst system is attributed interest and application in the pharmaceutical industry. The
to its enhanced basicity, and this is correlated against yield over Turner group at Manchester pioneered the use of engineered
a range of catalyst loadings. Numerous carboxylic acids and monoamine oxidases (MAO) for the preparation of chiral amines.
amines are reported to couple in good to excellent yields under The latest paper from this group describes the development
the optimized conditions, though aliphatic dicarboxylic and of a number of variants of MAOs derived from Aspergillus niger
B dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

having broad substrate scope and applications in deracemization in water at 100 °C, a reaction that produces only water and
reactions of substrates of pharmaceutical interest. The deracem- t-BuOH as waste. The reaction is tolerant of a number of func-
ization process consists of a cycle in which an MAO selectively tional groups and also is effective with substrates containing
oxidizes one enantiomer of a racemic amine to an imine, with pyridines and thiophenes. The authors propose a mechanism in
BH3−NH3 converting the resulting imine back to the racemic which the alcohol is first oxidized to the aldehyde, which then
amine. Repeated cycles thus provide a single enantiomer of the forms a hemiaminal with ammonia, followed by oxidation to
amine with high yield and ee. the primary amide. The drug piracetam was prepared using this
methodology as the final step in the synthesis (Green Chem.
2013, 15, 1956−1961).

Particularly noteworthy was the application of this method-


ology to the synthesis of (R)-harmicine using the oxidase
enzyme for two transformations. First, an indole intermediate
is nonselectively oxidized to an iminium ion which is trapped
by the nucleophilic indole to afford racemic harmicine. Next,
deracemization in the same reaction vessel via the oxidation/
reduction sequence using borane−ammonia produces (R)-
harmicine in overall 83% yield and >99% ee (J. Am. Chem. Soc. Elimination of highly toxic metals is a goal of green chemistry.
2013, 135, 10863−10869). However, when these metals are required for a particular
transformation, immobilization is an excellent way to minimize
exposure and permit reuse. Basavaraju et al. report the immobili-
zation of OsO4 by construction of a flow reactor with the
wall surface modified with a nanobrush polymer coating of
poly(4-vinylpyridine) used to immobilize OsO4 via ligation to
the pyridine units (Angew. Chem. 2013, 52, 6735−6738).

5. ASYMMETRIC HYDROGENATIONS
Liu et al. have developed an enantioselective metal-free hydro-
genation of imines by using simple chiral boranes as catalyst.
The active catalyst was generated in situ from hydroboration of
chiral dienes with HB(C6F5)2, thus allowing a wide range
Base metal catalysis is an area of renewed interest given the of catalysts to be evaluated. Once an optimal catalyst structure
inherent lack of sustainability for processes based on precious was identified, temperature and solvent effects were evaluated.
metals. Iron is a preferred metal catalyst, not only due to its Temperatures between 30 and 0 °C had little impact on enantio-
abundance but also for its low toxicity. Regarding oxidants, selectivity. Solvent effects were pronounced, and mesitylene was
H2O2 is a preferred green oxidant since the only waste identified as optimum. Using optimized conditions of 2.5 mol %
produced is water. Lenze and Bauer combine both these green diene and 5 mol % HB(C6F5)2 in mesitylene at 25 °C, a variety
aspects for the oxidation of 1,2-diols catalyzed by iron(II) of biaryl substituted imines were screened. Substitution was
complexes using H2O2 in MeCN, finding that the secondary limited primarily to the carbon-bound aryl group, and altera-
alcohol is selectively oxidized vs the primary alcohol. Five amine tion of the methyl group of the imine had little effect on
ligands were studied, with the symmetrical bis-pyridyl amine (1)
as the most effective. The authors conducted preliminary studies
to delineate the mechanism but did not draw firm conclusions.
Regardless of whether the reaction proceeds via hydride- or
hydrogen-atom abstraction, or via electron transfer, the stability
of the incipient secondary α-hydroxy carbon radical, cation, or
radical cation, over the primary species is likely to be responsible
for the observed selectivity. It is also worth noting that the
1,2-diol substrates do not suffer from C−C bond cleavage which
is the most common outcome with the majority of strong
oxidizing agents (Chem Commun. 2013, 49, 5889−5891).

Continuing with the theme of alcohol oxidations, Wu et al.


report a metal-free oxidation of primary and benzylic alcohols
to primary amides using tert-butylhydroperoxide and ammonia
C dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

enantioselectivity. Under mild reaction conditions asymmetric be effective, although with inferior yields and enantioselectivities
reduction is achieved up to 99% yield and 89% ee (J. Am. Chem. (Angew. Chem. Int. Ed. 2013, 52, 1−5).
Soc. 2013, 135, 6810−6813). An Ir(I)phosphine−phosphite complex has been developed
Continuous flow technology is considered to be one of and employed in the asymmetric reduction of various imine
the best alternatives to batch mode operations in R&D and hetereocyles, i.e. benzoxazines, benzoxazinones, benzothiazinones,
manufacturing. Duque et al. have developed a continuous-flow, and quinaxalinones for the first time. The catalyst loading was
solvent-free asymmetric hydrogenation strategy to reduce dibutyl found to be in the order of 0.5−2 mol %. This catalytic system
itaconate to (S)-dibutyl 2-methylsuccinate at room temperature works at room temperature and affords the product in excellent
under 5 bar H2 pressure in the presence of Rh−MeDuPHOS yields (94−99%) and enantioselectivities (94−99% ee). Mecha-
catalyst immobilized on alumina using phosphotungstic acid nistically, stepwise proton and hydride transfers are thought to
as the anchor. The conversion and enantioselectivity were found be the favored pathways to afford products (Org. Lett. 2013, 15,
to be up to 99% and 98%, respectively. The low operating 2066−2069).
pressure utilized in this protocol allows for a more efficient use
of hydrogen, and the product can be accessed without decom-
pression. Leaching of Rh is very low, allowing for extended run
times; however, the ee began to decrease after 23 h, and
conversion began to decrease after 47 h continuous reaction time
(Angew. Chem. Int. Ed. 2013, 52, 9805−9807).

The potential of BIPI (phosphinoimidazole) ligands in com-


bination with Ir(cod)BArF have been exploited in the asymmetric
reduction of difficult substrates, i.e. dimethylindene and
dimethyldihydronaphthalene. Screening identified ligand 2, having
the fluoride substituent, as extremely effective. A brief screen of
solvents was conducted, and dichloromethane gave higher conver-
sion compared to methanol or toluene. Various tetrasubstituted
olefin substrates have been tested, and the conversion and enantio-
Ru/SNNS [N,N-bis(2-tert-butylthiobenzylidene)-1,3- selectivities of these were found to be excellent under optimized
propanediamine] systems are analogous to chiral PNNP and conditions, consisting of 2 mol % catalyst, in dichloromethane at
salen Ru complexes and are found to have potential similar to 0 °C (Adv. Synth. Catal. 2013, 355, 1455−1463).
that of Ru diphosphine complexes. This is the first catalyst of its
kind that has been tested for asymmetric hydrogenation of
simple to hindered/unsaturated ketones in high yields (98%)
and enantioselectivity (95% ee). The substrate/base/catalyst
ratio was found to be in the order of 2000/100/1; in some
cases it was as high as 29300/1083/1. This system has better
features as it can tolerate air and moisture. Catalysts based on
Ru/SNNS show excellent chemoselectivity in the reduction of
unsaturated ketones and aldehydes with S/C ratios up to 106/1.
In situ formation of the catalyst without base was also found to

6. C−H ACTIVATION
Huang et al. have developed an iron-catalyzed oxidative radical
cross-coupling/cyclization to synthesize dihydrobenzofurans.
A combination of 10 mol % FeCl3 and 1.2 equiv DDQ are used to
couple electron-rich phenols and napthols with styrene or styrene
derivatives. The reaction is initiated through DDQ-promoted one-
electron phenol oxidation. Iron coordination of the oxidized
phenol stabilizes the carbon−centred radical which then under-
goes addition with styrene followed by C−O bond formation.
While the chemistry is quite mild and uses iron as a green catalyst,
one drawback is that 2 equiv of olefin are required to obtain high
yields (Angew. Chem. Int. Ed. 2013, 52, 7171−7155).
D dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

ionic strength of the reaction, a pH-dependent Brønsted acid-


cocatalyzed reaction was discovered where 1 equiv of TFA
provided the best yield. These conditions decrease reaction
time significantly to 5 h (23 without TFA) and improve yield to
90% (60% without TFA). Substitution on the amine−phenyl
A C−H lactonization method has been developed which ring was well tolerated as well as on the aromatic positions of
efficiently constructs biaryl lactones via carboxylic acid-directed the tetrahydroisoquinoline. A small sampling of other Michael
C−H functionalization/C−O bond reductive elimination. The acceptors also resulted in good yields. The authors present a
reaction uses a simple catalyst system, 5 mol % Pd(OAc)2 and significant amount of detail into the mechanistic understanding
15 mol % N-acetylated glycine, for C−H functionalization of of the reaction and studies to optimize the pH for this process,
2-aryl carboxylic acids. The terminal oxidant is (diacetoxyiodo)- highlighting the benefit this genre of Brønsted acid-cocatalyzed
benzene which is believed to generate a palladium(IV) inter- reaction systems may have on other radical processes (J. Org.
mediate that can more readily undergo C−O reductive elimination. Chem. 2013, 78, 4107−4114).
The reaction scope is broad; however, ortho-substituted biaryl
acids couple in moderate yields (∼50%) with >30% recovered
starting material. The authors demonstrate the utility of the
method through the synthesis of cannabinol which requires two
successive C−H lactonization steps (Org. Lett. 2013, 15,
2574−2577).

7. GREENER FLUORINATION
Mazzotti et al. reported a robust, scalable, and practical
palladium-catalyzed fluorination of arylboronic acids. Prior to
this work, the catalytic alternatives providing such function-
Pirnot et al. have developed a method for direct activation at alized aryl fluorides consisted of Pd-catalyzed fluorination of
the β-position to a carbonyl via photoredox activation. The aryl triflates, or the silver-catalyzed fluorination of stannanes.
system they optimized takes advantage of a common household The current work constitutes an obvious improvement from an
compact fluorescent bulb and a commercially available catalyst environmental standpoint. The Pd-catalyzed fluorination
with only 1 mol % loading. An amine cocatalyst facilitates uses terpyridyl Pd (II) complex as a precatalyst. It proceeds
enamine formation with the aldehyde, followed by enamine in DMF or acetonitrile and converts aryl trifluoroborates, or
oxidation and the subsequent nucleophilic radical coupling even their boronate esters or boronic acids, both electron−rich
with the arene at the β-position. Good yields from 44−86% for and poor, into their corresponding fluorides. The methodology
a variety of aldehydes and ketones in a cyclic 6-membered is compatible with a wide range of functional groups (ketones,
ring system when X = C and EWG = CN are reported. Also, amides, carboxylic acids, esters, bromides, heterocycles) and
a variety of electron-withdrawing groups including pyridine results in yields ranging from 63% to 99%. The paper comprises
systems where X = N delivers good yields, with the best yields useful mechanistic considerations which should further help in
provided by esters and sulfones at the ortho and para positions. optimization specific to a substrate in the future (J. Am. Chem.
The real highlight of this methodology may be the ability to use Soc. 2013, 135, 14012−14015).
a cinchona-based amine catalyst to deliver chiral products. At
this time they only highlight one example with cyclohexanone
delivering 50% ee, but studies are ongoing. The authors do
highlight the unfortunate formation of an equivalent of cyanide
during the course of the reaction, generating an aqueous waste
stream which must be treated accordingly. Finally, another draw-
back to consider is reaction time, with most reactions requiring
times in the 48 h range (Science. 2013, 339, 1593−1596).

Bloom et al. reported an elegant iron-catalyzed benzylic


fluorination method as a synthetic equivalent to 1,4-conjugate
addition of fluoride. Iron(II) acetylacetonate was demonstrated
to catalyze effectively the net conjugate fluorination formally
carried out with Selectfluor. While a large excess of fluoride
source is still being used, the ability to derivatize an activated
but unfunctionalized benzylic position offers an interesting
In another example of photocatylzed C−H activation, Yoon alternative in terms of feedstock. In addition, it enlarges the
and co-workers have developed a Brønsted acid-cocatalyzed scope and utility of iron as a catalyst. The methodology was
radical photocatalytic C−H activation at the α-amino site demonstrated on a wide range of aromatics and proceeded in
of tertrahydroisoquinolines. First, conditions were optimized modest yields (38% to 75%) with improved reactivity on esters
with the use of a 23 W compact fluorescent lamp and the and carboxylic acids. Although still suffering from modest yields,
commercially available Ru(bpy)3Cl2 catalyst. During the course the mild conditions combined with an operationally simple and
of investigations to improve the reaction rate by increasing the robust process makes this chemoselective transformation highly
E dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

attractive, and prone to future development (J. Org. Chem. 2013,


78, 11082−11086).

Terpenoid products obtained from the cyclisation of a


8. BIOCATALYSIS polyisoprene backbone have important applications from chiral
Many enzyme-catalysed reactions such as ketoreduction, building blocks in synthesis to therapeutic compounds such as
transamination, and monooxygenation are showing increasing taxol and artemisinin. Seitz et al. demonstrated that triterpene
use in chemical synthesis. Whilst the development of naturally cyclases can be used to generate di- and triheterocycles from
occurring reaction types is continuing, it is not necessary to functionalized, shortened polyisoprene units. This highlights
limit biocatalysis to reactions which have natural precedents. the versatility and flexibility of these enzymes to act as general
Coelho et al. demonstrated this concept by altering the activity Brønsted acid catalysts. Using a cyclase from Zymomonas mobilis,
of cytochrome P450BM3 from Bacillus megaterium from oxygena- a range of terminal nucleophiles could be used to produce
tion to cyclopropanation. By substituting a cysteine thiolate cyclic ethers, enol ethers, and lactones in reasonable yields
ligand with a weakly donating serine, the authors raised the (ChemBioChem 2013, 14, 436−439).
FeIII/II potential and allowed the enzyme to facilitate an
NAD(P)H-driven reaction. This mutation also removed mono-
oxygenation activity from the protein. The engineered enzyme
was shown to work in vivo for the cyclopropanation of styrene
under anaerobic conditions. Despite the use of ethyldiazoacetate
which is a drawback, the product was formed in good yield and
with high diastereo- and enantioselectivity (Nat. Chem. Biol.
2013, 9, 485−487).

Conducting more than one reaction in the same vessel can


make processes much more sustainable as it can reduce cycle
times and energy consumption, as well as minimise both yield
loss and waste. Enzymatic processes in particular have great
potential for one-pot cascades as reaction conditions tend to be
To further expand the transformations catalysed by the more similar than for chemical reactions. Liu and Li have
engineered P450BM3 enzyme, McIntosh et al. reported an intra- shown this approach is possible for the synthesis of (R)-2-alkyl-
molecular C−H amination. Using E. coli cells in vivo under δ-lactones from 2-alkylidenecyclopentanones. This system uses
anaerobic conditions, high yields, total turnover numbers (TTN), an enoate reductase from Acinetobacter sp. RS1 and a Baeyer−
and ee could be achieved. These results compare favorably with Villiger mono-oxygenase expressed in E. coli. The whole cells
those previously reported for this transformation by chemical expressing each enzyme are added sequentially to a single pot,
methods (Angew. Chem. Int. Ed. 2013, 52, 9309−9312). and the product lactones are produced in good yields and high
enantioselectivity (ACS Catal. 2013, 3, 908−911).

One of the more aspirational transformations in the ACS


Green Chemistry Institute Pharmaceutical Roundtable’s 2007
paper was the asymmetric hydrogentation of unfunctionalized
olefins, enamines, and imines. The reduction of this last class of
substrates has recently been reported by Rodriguez-Mata et al.
Using an (R)-selective imine reductase (IRED), Q1E1E0 from
Streptomyces kanamyceticus, 2-methyl-1-pyrroline could be
reduced to (R)-2-methylpyrrolidine in low yield but excellent ee.
Moreover, the crystal structure of the protein was solved, and
it is hoped that this may help deduce the mechanism of this
reaction and provide a basis for rational engineering of the Interestingly, the ee of the product is higher than that
enzyme (ChemBioChem 2013, 14, 1372−1379). observed in the initial reduction reaction, indicating possible
F dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

degradation of the product lactone. It was found that hydrolysis


of the (S)-2-alkyl-δ-lactone was being catalysed by Acinetobacter
sp. RS1, causing the ee of the (R)-δ-lactone to be upgraded.
As more and more novel biocatalysts and enzyme-catalysed
reactions are discovered, this brings into focus the need to
think differently about the way molecules could be constructed.
To assist in this regard, Turner and O’Reilly have proposed
that new guidelines for “biocatalytic retrosynthesis” should be
developed for future training and education of synthetic chemists
(Nat. Chem. Biol. 2013, 9, 285−288). substrate to generate a relatively stable tertiary cation. In the
absence of Et2O, dimerization of the substrate occurs (Angew.
9. REDUCTIONS Chem., Int. Ed. 2013, 52, 7492−7495).
It has long been known that hot, alkaline conditions decompose
sugars to hydrogen and a variety of low-molecular weight organic
species, e.g. lactate, formate, and acetate. Kumar et al. have
exploited this and used D-glucose as the sole reducing agent in a
catalyst-free process. There are some deficiencies in the system at
the present time. There is a requirement for an extended hold
at elevated temperatures, but at present, substrate concentra-
tions are only 0.25 M, and there is clear scope for optimisation. One crucial aspect to be considered for industrial application
Nonetheless, impressive chemoselectivity was reported with of FLPs (and indeed other novel research technologies) is the
halide, vinyl, carbonyl, and even additional nitro substituents tolerance towards impurities. Cost concerns typically mean that
remaining unaffected by the conditions. Yields were in the range solvents and other reagents are used at lower specification than
47−99% across a total of 22 substrates (RSC Adv. 2013, 3, in a research environment where prepurification is common. In
4894−4898). the FLP-catalyzed hydrogenation of imines, residual amounts of
aldehyde and water can lead to catalyst poisoning. Thompson
et al. have explored the use of scavengers with common FLP
systems and shown that small amounts of tri-isobutylaluminum
(TIBAL) or triethylsilane can prevent poisoning and can even
restore activity in a stalled reaction. Seven substrates were con-
sidered in this work, although chemoselectivity was not broadly
investigated apart from the presence of aryl halides. One aspect
of this work is that typical catalyst loadings are already around
0.5−2.5 mol % with the potential to reduce further although
loadings are higher for some electron-rich systems (Org. Process
Possibly the most studied of the metal−free reducing agents Res. Dev. 2013, 17, 1287−1292).
are the frustrated Lewis pairs (FLP). These systems deliver
both hydrogen activation and subsequent reduction of organic
compounds via noninteracting Lewis acid/base centres. A com-
mon component in such systems is tris(pentafluorophenyl)-
borane. Now Nicasio et al. have shown that related boranes
showing reduced Lewis acidity can also function effectively in
the reduction of electron-deficient olefins. They determined
that the rate-limiting step in such reductions was the transfer
of hydride from an intermediate borohydride species [e.g.,
(F5Ph)3BH−]. Moderating the Lewis acidity of the borane
species facilitated subsequent hydride transfer while still The above FLP cases might be viewed as niche examples that
allowing heterolytic hydrogen cleavage. Control experiments require lengthy reaction times and whose green credentials are
showed that both components were necessary for reaction to limited to avoiding the use of endangered metals. As the field
occur, applying the isolated borohydride reagent followed by develops, however, the scope for optimising Lewis acid/base
a DABCO·H+ quench was not successful. Optimal conditions partners and additives will increase and limitations of substrate
were reached using (2,4,6-F3Ph)3B as Lewis acid. A range of range and catalyst activity will reduce. We are beginning to see
substrates could be reduced using this system where the olefin that these approaches do show the potential for more tailored
was activated by ester, sulfonyl, and nitro groups. However, systems to be developed that offer scope for competitive metal-
apart from the nitroolefins, it was necessary for >1 electron- free reductions.
withdrawing group to be present (RSC Adv. 2013, 3, Base metal catalysis has gained in popularity as an alternative
21369−21372. Chem. Eur. J., 2013, 19, 11016−11020). to the use of precious metal catalysts. Typically having lower
Generally FLP systems utilize a P- or N-centred Lewis base, toxicity and greater abundance, this strategy is a green approach
Hounjet et al. have now shown that even simple ethers can to catalytic methodologies. Four papers that demonstrate the
function as promoters in certain situations and have demon- power of this strategy using iron catalysts were published during
strated the hydrogenation of 1,1-diphenylethylene with >95% the current review period. One strategy using in situ formed iron
conversion. It appears that the critical aspects of this approach heterocyclic carbene complexes was reported by Volkov et al.
are the Lewis basicity of the ether and the ability of the Using N,N-dimethylbenzamide as a model substrate, conditions
G dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

were refined that produced both high activity and selectivity sufficient to warrant strong consideration of this iron-catalyzed
for the amine, versus formation of the aldehyde and alcohol alternative of a very common transformation (Eur. J. Org. Chem.
products. Iron(II) acetate and the ionic liquid [Ph-HEMIM]- 2013, 2061−2065).
[OTf] were used under further refinement of silane and solvent.
Using these results, a mechanism was proposed that led to the
final optimized conditions. The reduction was reported with aryl
or heteroaryl tertiary amides, using 1 mol % iron(II) acetate,
1.1 mol % [Ph-HEMIM][OTf], 2.2 mol % n-BuLi, 1 mol %
LiCl, and 3 equiv polymethylhydrosiloxane (PMHS) in THF at
65 °C, affording the tertiary amine product in high yield with
relatively short reaction time. Reactions using primary or Another publication focussing on reduction of carboxylic
secondary amides failed to produce product (Eur. J. Org. Chem. esters and amides was reported by Fernández-Salas et al. In this
2013, 2066−2070). paper, the reduction of esters to alcohols was first approached
using silane-mediated reduction, catalyzed with potassium
hydroxide. After a screen of both silane and base catalyst,
conditions for the reduction of methyl benzoate were identified.
Using 1.1 equiv of phenylsilane with 4 mol % potassium
hydroxide under solvent-free conditions at room temperature
reduction reactions were complete within 1 h. A variety of esters
were reduced under these conditions, some required 60 °C to
Selective reduction of conjugated α,β-unsaturated aldehydes achieve complete reduction, and some required the use of a small
to provide allylic alcohols is a challenging problem in chemical amount of THF. Additional solvents were not screened, as the
synthesis. Wienhöfer et al. published a paper using iron catalysis authors appeared to prefer solvent-free conditions for this
to address this. Extending the use of an iron complex previously transformation. When applied to amides, these conditions
identified by the Beller group for hydrogenation of carbon provided the amines in high yield, though longer reaction times
dioxide and bicarbonate, they endeavoured to selectively reduce were required. Again, heating or the use of THF was required in
the model system of cinnamaldehyde. They quickly identified certain cases. Primary and secondary amides returned only
the need of an acidic cocatalyst and found that cinnamaldehyde starting materials; thus, this methodology is currently limited only
can be reduced selectively without reducing the olefin. A solvent to tertiary amides (Chem. Commun. 2013, 49, 9758−9760).
screen identified isopropyl alcohol as the optimum solvent, as
lower-chain alcohols lead predominately to formation of acetals.
High temperatures are required, the optimum being 140 °C, at
which temperature catalyst loading can be reduced to 0.05 mol %.
However, for practicality temperatures of 120 °C were chosen
for additional optimization. The optimum reaction conditions are
0.2 to 0.4 mol % [FeF(L)[BF4], 20 bar H2, 120 °C, trifluoroacetic
acid (2 mol %) in isopropyl alcohol. These conditions were
demonstrated on a number of conjugated aldehyde systems and
showed high functional group tolerance providing the allylic
alcohols in >95% yield in most cases. Additionally, a very nice
mechanistic investigation was conducted and reported in detail 10. ALCOHOL ACTIVATION FOR NUCLEOPHILIC
(Chem. Eur. J. 2013, 19, 7701−7707). DISPLACEMENT
Two groups have published studies into the N-alkylation of
heterocyclic amines with alcohols, in particular various benzazoles,
a common motif in pharmacophores. Bala et al. report on the
effectiveness of an iron catalyst for the transformation. Fe(II)-
phthalocyanine (FePc), an inexpensive catalyst the group has
developed for other chemistries, was found to provide good yield
(92%) in a model reaction of 2-aminobenzothiazole with benzyl
alcohol in toluene with sodium tert-butoxide at 100 °C; the
reaction did not proceed in the absence of catalyst. This protocol
Junge et al. published an iron-catalyzed reduction of carboxylic provides moderate to very good yields with a wide array of
ester to alcohols. After a brief screen of iron and ligand options, variously substituted heterocycles as well as providing various
an ideal combination was realized that used 5 mol % benzazoles via ring closures of ortho-substituted anilines reacting
Fe(stearate)2 with 10 mol % 1,2-diaminoethane. The reductant, with alcohols and 2 equiv base at 120 °C. The typical hydrogen-
polymethylhydrosiloxane (PMHS) was used in 3.0 equiv in transfer mechanism of alcohol oxidation, imine coupling, and
toluene at 100 °C. There was no mention of modifying the reduction was confirmed by intermediate trapping with modified
solvent for this transformation. Using these optimized conditions, reaction conditions (Green Chem 2013, 15, 1687−1693).
the reaction produced moderate and variable yields of the alcohol
products, and showed poor tolerance for other reducible
functionalities. Keto, nitro, and amide groups were fully reduced
under the reaction conditions, while the ester remained largely
intact. Nonetheless, the number of compatible functionalities is
H dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

In contrast, Donthiri et al. describe a metal-free N-alkylation Shimizu et al. prepared heterogeneous nickel nanoparticles
of 2-amino-substituted benzothiazoles, thiazoles, pyridine, supported on θ-alumina for the N-alkylation of anilines and
pyrazine, and pyrimidine with benzyl alcohols. The reaction some primary and secondary amines with a range of alcohols.
is conducted in toluene at 120 °C with 20 mol % sodium Reactions were conducted base free using 1 mol % of catalyst
hydroxide as catalyst. The reaction does not proceed in DMF, (5 wt % Ni on θ-alumina) in o-xylene at 144 °C, with 33 h−1
DMSO, glycerol, or water, whilst chlorobenzene gives a lower turnover frequency, comparing favorably to other heteroge-
yield in the model alkylation of 2-aminobenzothiazole with neous systems. Two mol % catalyst was used for the reaction of
4-chlorobenzyl alcohol. Weaker bases (Na2CO3, K2CO3) and aliphatic amines and alcohols. The catalyst was prepared by
lower temperatures (e.g., 100 °C) give lower yields, whereas reduction of supported nickel oxide under a flow of hydrogen
KOH, NaOtBu, and KOtBu are good (yield >85%), but not as at 500 °C prior to use. The catalyst was readily separated from
effective as NaOH. The yield of the reaction is independent the reaction mixture by centrifugation and could be reused at
of electron-donating or -withdrawing substituents in either least 3 times after regeneration under the reducing conditions.
the benzyl alcohol or 2-aminobenzothiazole with isolated yields Mechanistic studies support the hydrogen-borrowing mecha-
in the range 80−96%. N-benzyl 2-aminothiazole is obtained nism, and catalyst characterization studies led the authors to
in 96% isolated yield, comparable to Bala et al. at a higher conclude that the metal/support interface is probably the active
temperature without catalyst. The reaction only proceeds with site of the catalyst (ACS Catal. 2013, 3, 998−1005).
benzyl alcohols and pyridinemethanol (J. Org. Chem. 2013, 78,
6775−6781).

Continuing the heterogeneous catalyst theme, Yamada et al.


report the use of polyspiroborate iridium catalysts for the
alkylation of amines and ammonia in water. The insoluble
polymeric catalysts were prepared by treating a soluble linear
catecholborate polymer with an iridium species in a mixture of
The authors in both publications present studies to support DME−acetonitrile. Spectroscopic analysis of the resulting
in situ oxidation of the benzylalcohol to the corresponding insoluble product shows a material composed of anionic spirobo-
benzaldehyde as the first stage of the mechanism, facilitated by rate polymer with dicationic Ir2+, the iridium acting as a cross-
FePc in the first paper and oxygen in the second. These linker in the polymer. The catalyst (1 mol % Ir) was used for the
findings are consistent with the literature and may account for alkylation of anilines, benzylamine, dibenzylamine, morpholine,
the differences in reactivity highlighted here. and 4-hydroxymethylpiperidine with benzyl alcohols in water at
Heterogeneous catalysts are attractive for their ease of 100−150 °C for 24 h. Tribenzylamines could be prepared in
separation and potential for reuse; however, they can require 57−95% yield by heating aqueous ammonia with benzyl alcohol
extended reaction times at high temperature. Wang et al. have and the catalyst in pH 4 buffer at 150 °C in a microwave reactor
developed an N-heterocyclic carbene−iridium complex sup- for 24 h; ethyl acetate was used to extract the product during the
ported on a mesoporus silica (SBA-15) which was successfully work-up. The catalyst was removed by decantation and reused
used to catalyse the N-alkylation of anilines, benzylamine, three times (Synthesis 2013, 45, 2093−2100).
cyclohexylamine, and 2-aminopyrimidine and the β-alkylation
of 2° alcohols with benzyl alcohols (hexylalcohol was used in
one example of each transformation). The heterogeneous
catalyst was found to give a superior yield to the corresponding
homogeneous analogue in both transformations. The substrates
were mixed with the catalyst (1.5 mol % Ir), base and heated in
toluene at 110 °C. N-Alkylation reactions were run for 48 h
with NaHCO3 (0.5 equiv) and β-alkylation reactions 24 h with
KOH (1 equiv). The catalyst could be recovered and reused
8 times for β-alkylation and 12 times for amine alkylation. The Putra et al. have combined two hydrogen-transfer reactions
authors argue that the SBA-15 support plays an active role in for the enantioselective preparation of β-aminoalcohols from
the reactions by providing a water sink (Adv. Synth. Catal. 2013, racemic diols. Treatment of excess 1-phenyl-1,2-ethanediol with
355, 1117−1125). secondary amines in the presence of [RuCl2(p-cymene)]2 and
(S,R-Josiphos) for 24 h in toluene at 100 °C affords the amino
alcohol in good to high yield (99% by 1H NMR, 87% isolated
in 1 example) with moderate ee (up to 77%). Mechanistic
studies show that the diol is oxidised to the corresponding keto-
aldehyde which undergoes reductive amination of the aldehyde
followed by asymmetric hydrogenation of the ketone (Eur. J.
Org. Chem. 2013, 6146−6151) (Scheme 1).
Three publications describe the use of CO2 as a C1 building
block for amine methylation. Jacquet et al. utilise phenylsilane
(2 equiv) and a mixture of an N-heterocyclic carbene and
zinc chloride (5 mol %) in THF for N-methylation of
I dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

Scheme 1 great influence. Not surprisingly, electron-donating groups


enhance the reactivity over substrates bearing electron-with-
drawing groups. The mechanism is presumed to proceed via
addition of the 1,3-dicarbonyl compound to an N-acyliminium
ion intermediate (from the initial dehydration of the hydroxyl
group) followed by ring closure via Friedel−Crafts alkylation
(J. Heterocycl. Chem. 2013, 50, 501−505).

N-methylanilines. N,N-dimethylaniline was obtained in 95%


yield after 20 h at 100 °C, under 1 bar CO2. The influence of
electron-donating and -withdrawing substituents and their sub-
stitution position were explored, and the reaction was extended 12. CHEMISTRY IN WATER
to 1° amines and anilines with variable results; treatment of Chou and Raines describe the conversion of azides into diazo
aniline for 72 h afforded a mixture of N,N-dimethylaniline (79%) compounds in water. By the use of water as solvent the
and N-methyl-N-phenylformamide (21%) (Chem. Sci 2013, 4, applicability of this conversion in chemical biology can become
2127−2131). Li et al. have found a ruthenium-based catalyst of potential interest. The process is illustrated for the synthesis
system for use with phenylsilane (4 equiv) in toluene at 100 °C, of various diazoacetic acid esters and amides, and a wide range
under 30 bar CO2 for the N-methylation of a wider range of 2° of functional groups are tolerated.
amines; there are 22 examples with isolated yields ranging from
61−99% (Angew. Chem. Int. Ed. 2013, 52, 9568−9571).
Beydoun et al. use molecular hydrogen (60 bar) as the reductant
in the presence of a ruthenium catalyst in THF at 150 °C, under
20 bar CO2 (Angew. Chem. Int. Ed. 2013, 52, 9554−9557).
There is some overlap of examples with the previous paper and
neither method is superior in all cases. Anilines and primary
amines are dimethylated using both methods. All three groups
present studies that support a formamide intermediate.

The authors report the design and optimization of a water-


soluble activated phosphinoester used in stoichiometric
quantities to mediate the azide to diazo conversion in a phos-
phate buffer solution of pH 7. Especially, the reaction almost
exclusively produced the diazo compound without side-product
formation of Staudinger ligation if the conjugated acid of the
leaving group in the phosphinoester has a pKa < 7 (J. Am.
Chem. Soc. 2013, 135, 14936−14939).
By the developments in modern click chemistry the 1,2,3-
triazole group has become a privileged scaffold in many bioactive
molecules. An alternative preparation of these triazoles is by an
organocatalyzed 1,3-dipolar cycloaddition reaction of ketones
and azides. However, the reported reactions are restricted to the
use of organic solvents. Yeung et al. now describe the use of
water as solvent for this 1,3-dipolar cycloaddition using 20 mol %
of proline N,N-dioctylamide as organocatalyst. This reaction in
water at 80 °C is applied to a wide range of ketones and aromatic
azides (Green Chem. 2013, 15, 2384−2388).
11. FRIEDEL−CRAFTS CHEMISTRY
Wang et al. have demonstrated a one-pot synthesis of pyrrolo-
[1,2-a]quinolin-1-ones from the reactions of 5-hydroxy-1-
arylpyrrolidin-2-ones with 1,3-dicarbonyl compounds using
H3PO4/P2O5 or HOAc/H2SO4. The examples utilizing ethyl
acetoacetate and acetylacetone were generally high yielding;
however, when ethyl benzoyl acetate was employed, lower yields
were obtained. Presumably this is due to the steric hindrance of A new method for aroylation with CO and aryl iodides via
final cyclization. The substituents on the phenyl ring also have a ruthenium-catalyzed C−H activation of (hetero)arenes bearing
J dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

ortho-directing groups is described by Beller and co-workers. In another interesting article by Grego et al., details of a
Remarkably, the reactivity of this carbonylative C−C coupling catalytic fixed bed continuous process for the synthesis of
was dramatically increased by changing the organic solvent to methyl phenylcarbamate by carbamoylation of aniline by dimethyl
water. The best reaction conditions identified are 5 mol % of carbonate in the presence of zinc carbonate catalyst has been
RuCl2(cod) catalyst and 0.2 equiv of KOAc as additive, using reported. Detailed studies on parametric effects influencing this
NaHCO3 (2 equiv) as base and 30 bar CO at 120 °C for 20 h. phosgene-free process showed high conversions (∼98% at
Pyridine-2-yl, N-pyrazole, and pyrimidine-2-yl have been used 200 °C) and selectivity (97%) without any significant catalyst
as ortho-directing groups (Angew. Chem. Int. Ed. 2013, 52, deactivation even after 180 h despite a reaction stoppage and
6293−6297). restart in between (Org. Process Res. Dev. 2013, 17, 679−683).

Pedersen et al. made use of a setup where a continuous


stirred tank reactor (CSTR) and static mixer reactor were
The iron(III)-catalyzed direct nucleophilic substitution of connected in series to overcome challenges associated with
allylic (benzylic) alcohols with a wide range of nucleophiles has impurity formation and handling of suspended solids in plug
been described by Trillo et al. Although the application of flow reactors for the Grignard alkylation reaction of 2-chloro-
FeCl3 Lewis acid-catalyzed processes in water have been known thioxanthen-9-one with allylmagnesium chloride. The CSTR
for some time (e.g. in Mukiyama aldol reactions) examples of was used to generate and retain the solid phase, whereas an
direct substitution of free allylic alcohols in water are rare. outward line fitted with a filter was used to transfer the
Various sulfonamides and anilines, as well as benzyl carbamate, homogeneous liquid phase into plug flow static mixer reactor
benzamide, benzotriazole, and TMSN3 can be used as nitrogen to achieve higher conversions of the unreacted ketone (Org.
nucleophiles, but no conversion was observed with tert-butyl Process Res. Dev. 2013, 17, 1142−1148).
carbamate and basic amines (benzylamine, butylamine,
ammonia). Also carbon nucleophiles such as allyltrimethylsi-
lane, TMSCN, anisole, indole, or diethyl malonate can be used.
Use of FeO(OH)·2H2O instead of FeCl3 as the catalyst worked
equally well or even slightly better for less acidic sulfonamides,
anilines, and indole. The low toxicity of FeCl3, the use of water, Lehmann et al. carried out an inverse-electron-demand
and the lack of leaving group activation makes this conversion Kondrat’eva reaction to generate cycloalka[c]pyridines from
of special interest (ChemCatChem 2013, 5, 1538−1542). unactivated oxazoles and cycloalkenes in a continuous tubular
reactor under conditions of elevated temperature and pressure.
Continuous flow reactor characteristics offer much better
control and efficiency over batch reactors under the required
extreme operating conditions (230 °C and 750 psi), and the
process offers easy access to annulated pyridines (Org. Lett.
2013, 15, 3550−3553).
In recent years various successful examples of sp2 C−H bond
activation in water and the application in a variety of catalytic
cross-coupling reactions have been described in the literature.
These cross-coupling reactions and the role of water as a
solvent have been reviewed in a recent paper by Li and Dixneuf
(Chem. Soc. Rev. 2013, 42, 5744−5767).
Tomasek and Schatz review recent developments of olefin
metathesis in aqueous media. The two main strategies are
discussed: design and application of water-soluble catalysts and 14. GENERAL GREEN CHEMISTRY
the use of commercial catalysts under heterogeneous micellar Adams et al. from GSK have published an article on GSK’s
conditions (Green Chem. 2013, 15, 2317−2338). reagent guide which has the objective of embedding sustainabi-
lity into reagent selection. The assessment includes a combined
13. CONTINUOUS PROCESSING AND PROCESS environmental score, heath score, and safety score, a chemistry
INTENSIFICATION score which takes into account factors such as work-up, stoi-
Yoshida et al. published a featured article on the use of micro chiometry, and atom efficiency. Three of the reagent guides
reactors to enable flash chemistry (i.e., chemistry that is unsuit- are published in the paper with a further 12 shared in the
able for batch/semibatch processing) involving extremely fast Supporting Information. The format is impactful and easy to
reactions (<1 s) and highly reactive and unstable inter- use (Green Chem 2013, 15, 1542−1549).
mediates. The authors detail the approaches used supported Leahy et al., have published an article proposing seven key
with various examples of successful application of active in situ elements of an effective Green Chemistry program with examples
catalyst generation and multistep consecutive lithiation reac- from several pharmaceutical companies. The seven key elements
tions for protecting-group-free synthesis (Chem. Commun 2013, are (i) empowered Green Chemistry teams with management
49, 9896−9904). support, (ii) metrics and targets, (iii) resources and tools, (iv)
K dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Green Chemistry Highlights

education, (v) awareness and recognition, (vi) investments in


green technology, and (viii) external collaboration (Org. Process
Res. Dev. 2013, 17, 1099−1109).
Rakeshwar Bandichhor
Dr. Reddy’s Laboratories Ltd., Innovation Plaza, IPDO,
Bachupally, Hyderabad, Andhra Pradesh, India 500072
Apurba Bhattacharya
Dr. Reddy’s Laboratories Ltd., Innovation Plaza, IPDO,
Bachupally, Hyderabad, Andhra Pradesh, India 500072
Louis Diorazio
AstraZeneca, Macclesfield SK10 2NA, U.K.
Peter Dunn
Pfizer Global Research and Development, Ramsgate Road,
Sandwich, Kent, CT13 9NJ U. K.
Kenneth Fraunhoffer
Bristol-Myers Squibb, Co., One Squibb Drive, New
Brunswick, New Jersey 08903, United States
Fabrice Gallou
NovartisPharma AG, Forum 1, Novartis Campus, 4056
Basel, Switzerland
John Hayler*
GlaxoSmithKline, Stevenage, Hertfordshire, SG1 2NY
U.K.
Matthew Hickey*
Bristol-Myers Squibb, Co., One Squibb Drive, New
Brunswick, New Jersey 08903, United States
Bill Hinkley
GlaxoSmithKline, Research Triangle Park, North Carolina
27709, United States
David Hughes
Merck and Co., Inc., P.O. Box 2000, Rahway, New Jersey
07065, United States
Luke Humphreys
GlaxoSmithKline, Stevenage, Hertfordshire, SG1 2NY
U.K.
Bernard Kaptein
DSM Innovative Synthesis BV, P.O. Box 18, 6160
Maryland Geleen, The Netherlands
Suju Mathew
Pfizer Global Research and Development, Ramsgate Road,
Sandwich, Kent, CT13 9NJ U. K.
Lynnette Oh
GlaxoSmithKline, King of Prussia, Pennsylvania 19406,
United States
Paul Richardson
Pfizer Global Research and Development, 10578 Science
Center Drive, La Jolla, California 92121, United States
Timothy White
Eli Lilly, Indianapolis, Indiana United States
Stijn Wuyts
Janssen Pharmaceutical Companies of Johnson and
Johnson, Turnhoutseweg 30, B-2340 Beerse, Belgium

■ AUTHOR INFORMATION
Corresponding Authors
*E-mail: John.Hayler@gsk.com.
*E-mail: Matthew.Hickey@bms.com.

L dx.doi.org/10.1021/op500167y | Org. Process Res. Dev. XXXX, XXX, XXX−XXX

Вам также может понравиться